首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 593 毫秒
1.
 Li3N、Mg3N2和Ca3N2是高温高压下以hBN为原料合成cBN的催化剂。在实验中发现它们对常压高温下生成hBN的反应也表现出催化作用。对比了三种氮化物催化效果的差异,发现三种氮化物都只有在熔融状态下才能表现出催化效果,以及三种氮化物对生成hBN反应的催化效果与它们在高温高压下合成cBN反应的催化效果次序相似。提出了对合成hBN有催化作用的化合物也将对合成cBN表现出催化作用的观点。  相似文献   

2.
In this work we describe the results of Rutherford backscattering spectrometry, sheet resistivity measurements, X-ray diffractometry and conversion electron Mössbauer spectroscopy performed on thin film Fe-Al bilayered samples submitted to high vacuum furnace annealing. Isothermal anneals were performed at 570 K for time intervals ranging from 60 to 600 min. It is demonstrated that the diffusion of Al into Fe is smaller than the diffusion of Fe into Al for temperatures below 600 K. Sequential isochronous thermal anneals of 60 min were performed at temperatures ranging from 570 to 870 K, in order to study the stability of the formed phases. The stable Fe2Al5 intermetallic compound formed at 570 K decomposes at about 650 K, and the FeAl6 intermetallic compound appears at temperatures around 750 K.Work supported in part by CNPq and FINEP  相似文献   

3.
The solvation structure of magnesium, zinc(II), and alkaline earth metal ions in N,N‐dimethylformamide (DMF) and N,N‐dimethylacetamide (DMA), and their mixtures has been studied by means of Raman spectroscopy and DFT calculations. The solvation number is revealed to be 6, 7, 8, and 8 for Mg2+, Ca2+, Sr2+, and Ba2+, respectively, in both DMF and DMA. The δ (O C N) vibration of DMF shifts to a higher wavenumber upon binding to the metal ions and the shift Δν(= νbound − νfree) becomes larger, when the ionic radius of the metal ion becomes smaller. The ν (N CH3) vibration of DMA also shifts to a higher wavenumber upon binding to the metal ions. However, the shift Δν saturates for small ions, as well as the transition‐metal (II) ions, implying that steric congestion among solvent molecules takes place in the coordination sphere. It is also indicated that, despite the magnesium ion having practically the same ionic radius as the zinc(II) ion of six‐coordination, their solvation numbers in DMA are significantly different. DFT calculations for these metalsolvate clusters of varying solvation numbers revealed that not only solvent–solvent interaction through space but also the bonding nature of the metal ion plays an essential role in the steric congestion. The individual solvation number and the Raman shift Δν in DMF–DMA mixtures indicate that steric congestion is significant for the magnesium ion, but not appreciable for calcium, strontium, and barium ions, despite the solvation number of these metal ions being large. Copyright © 2006 John Wiley & Sons, Ltd.  相似文献   

4.
5.
Cluster surface collisions of C+ 60 (Buckminsterfullerene) ions with pyrolytic graphite are studied by time of flight mass spectroscopy in the energy range of 100 eV to 1500 eV at oblique angles of incidence extending our previous studies with steep angles of incidence. Analysis of the velocity dependence of the scattered cluster ions reveals that the normal and tangential component of the ion velocity have different significance for the collision dynamics. The tangential component is partially transformed into rotational energy of the C+ 60 during the interaction with the surface as may be explained by an intuitive and astonishingly simple rolling ball model. In contrast, the normal component of the velocity appears to determine the amount of energy transferred into vibrational and deformational energy of projectile and target. At high normal energies the C+ 60 disintegrate significantly and can also be partially deposited onto the surface.  相似文献   

6.
在气体样品池条件下,研究了Rb(5PJ) (He、N2)碰撞能量转移过程.调频半导体激光器稍微调离共振线,激发Rb原子至Rb(5P3/2)态,在不同的He或N2气压下,测量了直接5P3/2→5S1/2荧光和转移5P1/2→5S1/2荧光.对于5PJ与He的碰撞.电子态能量仅能转移为He原子的平动能.在与N2的碰撞中,向分子振转态的转移是重要的.利用速率方程分析,可以得到碰撞转移速率系数,对于He,5P3/2→5S1/2转移速率系数为2.23×10-12cm3s-1.对于N2,测量5PJ He和5PJ N2二种情况下荧光的相对强度比,利用最小二乘法确定5P3/2→5S1/2转移速率系数为4.38×10-11cm3s-1,5PJ态猝灭速率系数为5.45×10-11cm3s-1.与其他实验结果进行了比较.  相似文献   

7.
《光谱学快报》2013,46(4-5):521-537
Abstract

Proton and carbon‐13 NMR data are presented for 5‐methoxytryptamine, 1; 6‐methoxytryptamine, 2; N,N‐diisopropyl‐5‐methoxytryptamine, 3, (5‐MeO‐DIPT); and N,N‐diisopropyl‐5‐methoxyindole‐3‐glyoxylamide, 4, at 300 MHz (1H) and 75 MHz (13C) in CDCl3 at ambient temperature. Compound 3, considered a potential hallucinogen, had been placed into Schedule I of the Controlled Substances Act, effective April 4, 2003, by the U.S. Drug Enforcement Administration. Compound 4 can serve as a possible precursor to 3. We believe that these are the first proton NMR assignments obtained at medium field (7 tesla) using selective homodecoupling and two‐dimensional homonuclear chemical shift correlation spectra (using one or more of the COSY45, COSY90, and COSYLR experiments) for rigorous aryl proton assignments in this group of compounds. Significant observed differences in the proton and carbon‐13 NMR spectra should allow facile distinction of the 5‐methoxy series, 1 and 3, from the 6‐methoxy series, 2. Energy minimizations to obtain optimized structures for each compound were performed at the Hartree–Fock level with the 6‐31G* basis set, and the resulting geometries are discussed. The presented geometry calculations appear to be the most accurate reported to date for 1 based on the basis set employed, and the first HF/6‐31G* structures for compounds 2, 3, or 4. Appreciable geometry differences in 3 and 4 for the pendant sidechain containing the N[CH(CH3)2]2 moiety are noteworthy. Proximity of the carbonyl oxygens in 4 to H2 and H4 is suggested as a possible contributing factor in the deshielding of these protons in the NMR spectrum.  相似文献   

8.
Two series of new N‐1 acylindazoles containing 5‐ or 6‐nitro groups were synthesized with moderate to good yields and characterized by IR and NMR spectroscopy. Cyclic voltammetry in aprotic media was utilized for the electrochemical characterization of the compounds. The calculated reduction potentials in physiological conditions are similar to those of known commercial antichagasic drugs. Therefore, the novel series reported herein are prospective candidates for antichagasic biological evaluation. Theoretical calculation results indicate that the studied dinitro derivatives undergo a single step reduction process because of the energy proximity of their radical and anionic state. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

9.
Abstract

Mid-infrared, far-infrared, and Raman vibrational spectroscopic studies were combined with density functional theory (DFT) calculations and normal coordinate force field analyses for N,N′-dimethylurea (DMU), N,N,N′,N′-tetramethylurea (TMU), and N,N′-dimethylpropyleneurea (DMPU: IUPAC name 1,3-dimethyltetrahydropyrimidin-2(1H)-one). The equilibrium molecular geometry of DMU (all three conformers), TMU, and DMPU and the frequencies, intensities, and depolarization ratios of their fundamental infrared (IR) and Raman vibrational transitions were obtained by DFT calculations. The vibrational spectra were fully analyzed by normal coordinate methods as well. A starting force field for DMPU was obtained by adapting corresponding force constants for DMU and TMU, resulting after refinements in the stretching force constants C=O (7.69, 7.30, 7.68 N·cm?1), C–N (5.16, 5.55, 5.05 N·cm?1), and C-Me (5.93, 4.00, 4.22 N·cm?1) for DMU, TMU, and DMPU, respectively. The dominating conformer of liquid DMU was identified as trans-trans, strong intermolecular hydrogen bonding was verified in solid DMU, and weak dipole–dipole association was found in liquid TMU and in DMPU. Special attention was paid to analyzing the methyl group frequencies, which revealed deviations from local C3v symmetry. A linear correlation was found between the CH stretching force constants and the inverse of the CH bond lengths (1/r 2). The averaged NH stretching frequencies of gaseous, dissolved, and solid urea and of DMU, with variations for hydrogen bonding of different strength, are linearly correlated to the NH stretching force constants. Characteristic skeletal vibrations were assigned for a broad variety of urea derivatives and also for pyrimidine derivatives, which all contain the N2C=O entity. The very strong IR bands of C=O stretching (1,676 ± 40 cm?1) and asymmetric CN2 stretching (1,478 ± 60 cm?1), and the very intense Raman feature of symmetric CN2 stretching or ring breathing (757 ± 80 cm?1), can be recognized as fingerprint bands also for the pyrimidine derivatives cytosine, thymine, and uracil, which all are nucleobases in DNA and RNA nucleotides.  相似文献   

10.
Raman spectra of acetic acid (AA), N,N‐dimethyl formamide (DMF) and their binary mixtures with varying mole fraction of the AA were recorded in the region 300–1750 cm−1 to investigate the formation of self‐associated dimer and hydrogen‐bonded complexes in a mixed system. The observed spectral features of the CO stretching mode suggest the formation of self‐association with a smaller aggregation size, and also indicate the presence of repulsive interactions between AA and DMF. The existence of two kinds of AA molecules (free and complex) is elucidated from the splitting of the OC O deformation mode. The intermolecular hydrogen‐bond formation and the possibility of attractive interaction between AA and DMF are also examined from the observed spectral features in the CCO symmetric stretching mode of AA, and CN symmetric stretching mode of DMF. Copyright © 2006 John Wiley & Sons, Ltd.  相似文献   

11.
Recent studies have shown that general‐base assisted catalysis is a viable mechanistic pathway for hydrolysis of smaller anhydrides. Therefore, it is the central purpose of the present work to compare and contrast the number of hydrogen atoms in‐flight and stationary in the transition state structure of the base‐catalyzed mechanisms of 2 hydrolytic reactions as well as determine if any solvent effects occur on the mechanisms. The present research focuses on the hydrolytic mechanisms of N,N‐dimethylformamide (DMF) and acetic anhydride in alkali media of varying deuterium oxide mole fractions. Acetic anhydride has been included in this study to enable comparisons with DMF hydrolysis. Comparative studies may give synergistic insight into the detailed structural features of the activated complexes for both systems. Hydrolysis reactions in varying deuterium oxide mole fractions were conducted in concentrations of 2.0M , 2.5M , and 3.0M for DMF and 0.10M for acetic anhydride at 25°C. Studies in varying deuterium mole fractions allow for proton inventory analysis, which sheds light on the number and types of hydrogen atoms involved in the activated complex. For these systems, this type of study can distinguish between direct nucleophilic attack of the hydroxide ion on the carbonyl center and general‐base catalysis by the hydroxide ion to facilitate a water molecule attacking the carbonyl center. The numerical data are used to discuss 3 possible mechanisms in the hydrolysis of DMF.  相似文献   

12.
Held has proposed a coordinate- and gauge-free integration procedure within the ghp formalism built around four functionally independent zero-weighted scalars constructed from the spin coefficients and the Riemann tensor components. Unfortunately, a spacetime with Killing vectors (and hence cyclic coordinates in the metric, and in all quantities constructed from the metric) may be unable to supply the full quota of four scalars of this type. However, for such a spacetime additional scalars may be supplied by the components of the Killing vectors. As an illustration we investigate the vacuum type N spaces admitting a Killing vector and a homothetic Killing vector. In a direct manner, we reduce the problem to a pair of ordinary differential operator master equations, making use of a new zero-weighted ghp operator. In two different ways, we show how these master equations can be reduced to one real third-order operator differential equation for a complex function of a real variable—but still with the freedom to choose explicitly our fourth coordinate. It is then easy to see there is a whole class of coordinate choices where the problem reduces essentially to one real third-order differential equation for a real function of a real variable. It is also outlined how the various other differential equations, which have been derived previously in work on this problem, can be deduced from our master equations.  相似文献   

13.
Densities ρ, viscosities η, and refractive indices nD, of the binary and ternary mixtures formed by cyclohexanone + N,N-dimethylacetamide + N,N-diethylethanolamine were measured at (298.15, 308.15, and 318.15) K for the liquid region and at ambient pressure for the whole composition ranges. The excess molar volumes VmE, viscosity deviations Δη, and refractive index deviations ΔnD, were calculated from experimental densities and refractive indices. The excess molar volumes are positive over the mole fraction range for binary mixtures of cyclohexanone(1) + N,N-dimethylacetamide (2) and N,N-dimethylactamide (2) + N,N-diethylethanolamine (3) and increase with increasing temperatures from (298.15 to 318.15) K. The excess molar volumes of cyclohexanone (1) + N,N-diethylethanolamine (3) are S-shaped dependence on composition with negative values in the N,N-diethylethanolamine rich-region and positive values at the opposite extreme and increase with increasing temperatures from (298.15 to 318.15) K. The excess molar volumes are positive over the whole mole fraction ranges for the ternary mixtures at all temperatures. Viscosity deviations are negative over the mole fraction range for all binary and ternary mixtures and decrease with increasing temperatures from (298.15 to 318.15) K. Refractive index deviations are negative over the mole fraction range for all binary and ternary mixtures and increase with increasing temperatures from (298.15 to 318.15) K. The experimental data of constitute were correlated as a function of the mole fraction by using the Redlich–Kister equation for binary and , Cibulka, Jasinski and Malanowski , Singe et al., Pintos et al., Calvo et al., Kohler, and Jacob–Fitzner for ternary mixture, respectively. McAllister's three body, Hind, and Nissan–Grunberg models were used for correlating the kinematic and dynamic viscosity of binary mixtures. The experimental data of the constitute binaries are analyzed to discuss the nature and strength of intermolecular interactions in these mixtures.  相似文献   

14.
Neha Katyal  Amitava Roy 《Optik》2011,122(3):207-210
Two-wave mixing (TWM) in transmission geometry of photorefractive (PR) crystals of 23 symmetry such as BSO, BGO is investigated by solving a set of coupled wave equations and taking their optical activity as one of the effective parameter. Coupled wave equations are solved numerically by fourth order Runge-kutta method and results are presented in graphical form. The effects of various parameters such as crystal thickness, optical activity, coupling coefficient, absorption, and input polarization angle of pump beam has been studied on the input beam polarization components. It has been observed that by varying the behaviour of pump beam polarization, signal beam can be modified. This can be used in various applications such as spatial light modulation and optical information processing, etc.  相似文献   

15.
通过Al取代B在B12N12中掺杂,基于密度泛函理论对B6Al6N12的稳定性进行了系统研究.表明在B6Al6N12中,Al和B倾向于相对分开.在最稳定结构中,B原子和Al原子分别处在笼的两半.我们还分析了B12-nAlnN12 (n=0, 6, 12)的电子结构,B12N12的能隙为6.84eV,掺Al后其能隙明显变窄,Al12N12和B6Al6N12的能隙分别为3.91eV和4.08eV.NBO分析表明,B12N12中B/N原子的电荷分别为±1.17,Al12N12中Al/N 原子的电荷分别为±1.85.在B6Al6N12中,B/Al原子上的电荷分别为1.06~1.12和1.86~1.91,N原子上的电荷为-1.18~-1.83.  相似文献   

16.
通过Al取代B在B12N12中掺杂,基于密度泛函理论对B6Al6N12的稳定性进行了系统研究.表明在B6Al6N12中,Al和B倾向于相对分开.在最稳定结构中,B原子和Al原子分别处在笼的两半.我们还分析了B12-nAlnN12 (n=0, 6, 12)的电子结构,B12N12的能隙为6.84eV,掺Al后其能隙明显变窄,Al12N12和B6Al6N12的能隙分别为3.91eV和4.08eV.NBO分析表明,B12N12中B/N原子的电荷分别为±1.17,Al12N12中Al/N 原子的电荷分别为±1.85.在B6Al6N12中,B/Al原子上的电荷分别为1.06~1.12和1.86~1.91,N原子上的电荷为-1.18~-1.83.  相似文献   

17.
红外光谱在线监控对二苯脲合成反应进程的研究   总被引:1,自引:0,他引:1  
利用ReactIRTM4000实验室在线红外分析系统对脲和苯胺合成二苯脲反应进行在线监控,通过监测反应物脲(1 420 cm-1)和苯胺(1 270 cm-1)、中间产物苯基脲(1 339 cm-1)及产物二苯脲(1 312 cm-1)特征峰的变化,推断出反应体系中各组分浓度的变化,从而清楚地了解了该反应的进程。研究的结果证明脲和苯胺合成二苯脲的过程是脲和苯胺先生成中间产物苯基脲,苯基脲再和苯胺生成二苯脲。结果表明在线红外分析能够快速、简便、准确地确定出反应的最佳时间,从而达到对合成反应的进程进行控制的目的。  相似文献   

18.
在pH =5 7混合酸 氢氧化钠缓冲介质中 ,研究了铬 (Ⅵ )催化过氧化氢氧化对 氨基二甲替苯胺盐酸盐和H酸 钠盐的氧化偶联反应及其机理 ,建立了催化光度法测定痕量铬 (Ⅵ )的新方法。该方法线性范围为0 0 8~ 0 4 0mg·L-1 ,检出限为 2 0× 10 -6g·L-1 ,相对标准偏差为 1 1% ,直接用于废水中铬 (Ⅵ )的测定 ,结果满意  相似文献   

19.
The first N‐allenyl derivative of trifluoromethanesulfonamide, N‐benzyl‐N‐(allenyl)trifluoromethanesulfonamide ( 1 ), was studied experimentally by the FT‐IR spectroscopy and theoretically at the DFT and MP2 levels of theory. The intramolecular interaction of the nitrogen atom with the triflyl and the allenyl group was studied in comparison with the analogously substituted vinyl derivatives. Compound 1 in heptane solution at 295–183 K exists as an equilibrium mixture of conformational isomers. Protonation at different basic sites in a series of reference molecules is studied theoretically. The central C2 atom of the allenyl group in 1 has the highest proton affinity, which is 16 kcal/mol higher than in the N‐vinyl analogues. The relative ability of the allenyl and vinyl groups to conjugation with an electron‐rich and electron‐deficient nitrogen atom lone electron pair is discussed. From the NBO analysis, the conjugation of the nitrogen lone electron pair with the allenyl group is much stronger than with the vinyl group. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

20.
By ab-initio calculation we show that the (Ga,Fe)N ground state may be changed from anti-ferromagnetic to ferromagnetic by acceptor defect like Ga vacancies. The electronic structures are calculated by using the Korringa-Kohn-Rostoker (KKR) method combined with coherent potential approximation (CPA). We show that we can increase the magnetic moment of Fe in p-type GaN by oxygen co-doping. Mechanism of exchange interactions between magnetic ions in p-type (Ga,Fe)N is also studied. The effect of external magnetic field on the electronic structure of (Ga, Fe)N and p-type (Ga, Fe)N is investigated.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号